This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Hyperbolic L-space knots and their Upsilon invariants

Masakazu Teragaito International Institute for Sustainability with Knotted Chiral Meta Matter (WPI-SKCM2), Hiroshima University, 1-3-1, Kagamiyama, Higashi-hiroshima 7398526, Japan [email protected]
Abstract.

For a knot in the 33–sphere, the Upsilon invariant is a piecewise linear function defined on the interval [0,2][0,2]. For an L–space knot, the Upsilon invariant is determined only by the Alexander polynomial of the knot. We exhibit infinitely many pairs of hyperbolic L–space knots such that two knots of each pair have distinct Alexander polynomials, so they are not concordant, but share the same Upsilon invariant. Conversely, we examine the restorability of the Alexander polynomial of an L–space knot from the Upsilon invariant through the Legendre–Fenchel transformation.

2020 Mathematics Subject Classification:
Primary 57K10; Secondary 57K18
The author has been supported by JSPS KAKENHI Grant Number 20K03587.

1. Introduction

For a knot KK in the 33–sphere S3S^{3}, Ozsváth, Stipsicz and Szabó [30] defined the Upsilon invariant ΥK(t)\Upsilon_{K}(t), which is a piecewise linear real-valued function defined on the interval [0,2][0,2]. This invariant is additive under connected sum of knots, and the sign changes for the mirror image of a knot. Also, it gives a lower bound for the genus, the concordance genus and the four genus. Although it is originally defined through some modified knot Floer complex, Livingston [23] later gives an alternative interpretation on the full knot Floer complex CFK\mathrm{CFK}^{\infty}.

As the most important feature, the Upsilon invariant is a concordance invariant, so it is obviously not strong to distinguish knots, although it has been used to establish various powerful results about independent elements in the knot concordance group [11, 16, 30, 38]. For a smoothly slice knot, the Upsilon invariant is the zero function. It depends only on the signature for an alternating knot or a quasi-alternating knot [30]. Also, it is determined by the τ\tau–invariant for concordance genus one knots [11].

In this paper, we concentrate on L–space knots, which are recognized to form an important class of knots in recent research. A knot is called an L–space knot if it admits a positive surgery yielding an L–space. Positive torus knots are typical examples of L–space knots. Note that any non-trivial L–space knot is prime [19] and non-slice [27]. For an L–space knot, the Upsilon invariant is determined only by the Alexander polynomial [30, Theorem 6.2].

There is another interesting route to lead to the Upsilon invariant of an L–space knot. The Alexander polynomial gives the formal semigroup [37], in turn, the gap function [7]. These notions have the same information as the Alexander polynomial. Then the Upsilon invariant is obtained as the Legendre–Fenchel transform of the gap function [6].

In general, the gap function for an L–space knot is not convex, so the further Legendre–Fenchel transformation on the Upsilon invariant does not return the original gap function. Thus there is a possibility that distinct gap functions, equivalently Alexander polynomials, correspond to the same Upsilon invariant. In other words, it is expected to exist non-concordant L–space knots with the same Upsilon invariant. We remark that the Alexander polynomial is a concordance invariant for L–space knots [19]. Our main result shows that this is possible among hyperbolic L–space knots.

Theorem 1.1.

There exist infinitely many pairs of hyperbolic L–space knots K1K_{1} and K2K_{2} such that they have distinct Alexander polynomials but share the same non-zero Upsilon invariant.

Thus two hyperbolic L–space knots in our pair are not concordant. In the literature, there are plenty of examples of non-concordant knots sharing the same Upsilon invariant [1, 11, 17, 38, 39, 40, 42]. However, they use either connected sums of torus knots or satellite knots, which are not hyperbolic.

Since the Upsilon invariant is determined only by the Alexander polynomial for an L–space knot, any pair of L–space knots sharing the same Alexander polynomial have the same Upsilon invariant. For example, the hyperbolic L–space knot t09847 in the SnapPy census has the same Alexander polynomial as the (2,7)(2,7)–cable of T(2,5)T(2,5), which is an L–space knot. There are infinitely many such pairs consisting of a hyperbolic L–space knot and an iterated torus L–space knot (found in [3]).

However, we checked Dunfield’s list of 632632 hyperbolic L–space knots ([2, 3]), and confirmed that there is no duplication among their Alexander polynomials and that there is no one sharing the same Alexander polynomial as a torus knot. This leads us to pose a question.

Question 1.2.
  1. (1)

    Do there exist hyperbolic L–space knots which have the same Alexander polynomial, equivalently, which are concordant?

  2. (2)

    Does there exist a hyperbolic L–space knot which is concordant to a torus knot?

In general, it is rare that the Alexander polynomial of an L–space knot is restorable from the Upsilon invariant. The reason is the fact that the gap function, which has the same information as the Alexander polynomial, is not convex, and the Upsilon invariant depends only on the convex hull of the gap function. In fact, our knots in Theorem 1.1 are designed so that they have distinct Alexander polynomials, but their gap functions share the same convex hull, so the same Upsilon invariant.

On the other hand, there is a chance that the gap function is restorable from its convex hull. This means that the Alexander polynomial is also restorable from the Upsilon invariant through the Legendre–Fenchel transformation. We can give infinitely many such gap functions, equivalently Alexander polynomials, but there lies a hard question, called a geography question, whether such gap function can be realized by an L-space knot or not.

In this paper, we can give only two hyperbolic L–space knots whose Alexander polynomials are restorable from the Upsilon invariants.

Theorem 1.3.

Let KK be the hyperbolic L–space knot t09847 or v2871 in the SnapPy census. Then the Alexander polynomial ΔK(t)\Delta_{K}(t) of KK is restorable from the Upsilon invariant ΥK(t)\Upsilon_{K}(t). That is, the equation ΥK(t)=ΥK(t)\Upsilon_{K}(t)=\Upsilon_{K^{\prime}}(t) implies ΔK(t)=ΔK(t)\Delta_{K}(t)=\Delta_{K^{\prime}}(t) (up to units) for any other L–space knot KK^{\prime}.

In Section 2, we give a pair of knots K1K_{1} and K2K_{2}, which yields an infinite family of pairs of L–space knots. In Section 3, we calculate their Alexander polynomials and the formal semigroups, which are sufficient to prove that the knots are hyperbolic. Section 4 gives the gap functions and their convex hulls, and confirm that they correspond to the same Upsilon invariant. Section 5 shows that the knots admit L–space surgery through the Montesinos trick, which completes the proof of Theorem 1.1. In the last section, we investigate the restorability of Alexander polynomial from the Upsilon invariant, and prove Theorem 1.3.

2. The pairs of hyperbolic L-space knots

For any integer integer n1n\geq 1, the surgery diagrams illustrated in Figure 1 define our knots K1K_{1} and K2K_{2}, where the surgery coefficient on C1C_{1} is 1/n-1/n and that on C2C_{2} is 1/2-1/2. The images of KK after these surgeries in (1) and (2) of Figure 1 give K1K_{1} and K2K_{2}, respectively. (The link with orientations is placed in a strongly invertible position, and the axis is depicted there for later use.)

Refer to caption
Figure 1. The knots K1K_{1} and K2K_{2} are the images of KK after performing (1/n)(-1/n)–surgery on C1C_{1} and (1/2)(-1/2)–surgery on C2C_{2}.

Hence, our knots are the closures of 44–braids

(σ2σ1σ3σ2)(σ1σ2σ3)4nσ21(σ2σ3)6and(σ2σ1σ3σ2)(σ1σ2σ3)4nσ31(σ2σ3)6,(\sigma_{2}\sigma_{1}\sigma_{3}\sigma_{2})(\sigma_{1}\sigma_{2}\sigma_{3})^{4n}\sigma_{2}^{-1}(\sigma_{2}\sigma_{3})^{6}\quad\text{and}\quad(\sigma_{2}\sigma_{1}\sigma_{3}\sigma_{2})(\sigma_{1}\sigma_{2}\sigma_{3})^{4n}\sigma_{3}^{-1}(\sigma_{2}\sigma_{3})^{6},

where σi\sigma_{i} is the standard generator of the 44–strand braid group. When n=1n=1, K1K_{1} is m240, and K2K_{2} is t10496 in the SnapPy census [9].

Theorem 1.1 immediately follows from the next.

Theorem 2.1.

For each integer n1n\geq 1, the knots K1K_{1} and K2K_{2} defined above satisfy the following.

  • (1)

    They are hyperbolic.

  • (2)

    (16n+21)(16n+21)–surgery on K1K_{1} and (16n+20)(16n+20)–surgery on K2K_{2} yield L–spaces.

  • (3)

    Their Alexander polynomials are distinct.

  • (4)

    They share the same Upsilon invariant.

Proof.

This follows from Lemmas 3.6, 5.1, 5.2, Theorems 3.1, 3.2, and Corollary 4.3. (To see that the Alexander polynomials of K1K_{1} and K2K_{2} are distinct, it may be easier to compare their formal semigroups. From Propositions 3.3 and 3.4, we have that 4n+7𝒮K14n+7\in\mathcal{S}_{K_{1}}, but 4n+7𝒮K24n+7\not\in\mathcal{S}_{K_{2}}.) ∎

Each diagram in Figure 1 has a single negative crossing, but it can be cancelled obviously with some positive crossing. Hence both knots are represented as the closures of positive braids, which implies that they are fibered [33]. Then it is straightforward to calculate their genera g(Ki)g(K_{i}), and we see that g(K1)=g(K2)=6n+6g(K_{1})=g(K_{2})=6n+6.

Also, if once we know that KiK_{i} is an L–space knot, then rr–surgery on KiK_{i} gives an L–space if and only if r2g(Ki)1=12n+11r\geq 2g(K_{i})-1=12n+11 by [14, 29]. Our choices of surgery coefficients in Theorem 2.1(2) come from the manageability in the process of the Montesinos trick in Section 5.

3. Alexander polynomials

We calculate the Alexander polynomials of K1K_{1} and K2K_{2}. Since K1K_{1} and K2K_{2} are obtained from KK by performing some surgeries on C1C_{1} and C2C_{2}, we mimic the technique of [3].

Theorem 3.1.

The Alexander polynomial of K1K_{1} is given as

ΔK1(t)=i=0n(t8n+12+4it8n+11+4i)+(t8n+9t8n+8)+i=0n(t4n+6+4it4n+4+4i)+(t4n+3t4n+1)+i=0n1(t4+4it1+4i)+1.\begin{split}\Delta_{K_{1}}(t)&=\sum_{i=0}^{n}(t^{8n+12+4i}-t^{8n+11+4i})+(t^{8n+9}-t^{8n+8})+\sum_{i=0}^{n}(t^{4n+6+4i}-t^{4n+4+4i})\\ &\quad+(t^{4n+3}-t^{4n+1})+\sum_{i=0}^{n-1}(t^{4+4i}-t^{1+4i})+1.\end{split}
Proof.

Let L=KC1C2L=K\cup C_{1}\cup C_{2} be the oriented link illustrated in Figure 1(1). Its multivariable Alexander polynomial is

ΔL(x,y,z)=x6y3z2+x5y2zx3y3z2+x3y2z2x3y2zx2y2z2+x4y+x3yzx3y+x3xyz1,\begin{split}\Delta_{L}(x,y,z)&=x^{6}y^{3}z^{2}+x^{5}y^{2}z-x^{3}y^{3}z^{2}+x^{3}y^{2}z^{2}-x^{3}y^{2}z\\ &\quad-x^{2}y^{2}z^{2}+x^{4}y+x^{3}yz-x^{3}y+x^{3}-xyz-1,\end{split}

where the variables x,y,zx,y,z correspond to the (oriented) meridians of KK, C1C_{1}, C2C_{2}, respectively. (We used [9, 18] for the calculation.)

Performing (1/n)(-1/n)–surgery on C1C_{1} and (1/2)(-1/2)–surgery on C2C_{2} changes the link KC1C2K\cup C_{1}\cup C_{2} to K1C1nC2nK_{1}\cup C_{1}^{n}\cup C_{2}^{n}. These two links have homeomorphic exteriors. Hence the induced isomorphism of the homeomorphism on their homology groups relates the Alexander polynomials of two links [12, 25].

Let μK\mu_{K}, μC1\mu_{C_{1}} and μC2\mu_{C_{2}} be the homology classes of meridians of KK, C1C_{1}, C2C_{2}, respectively. We assume that each meridian has linking number one with the corresponding component. Furthermore, let λK\lambda_{K}, λC1\lambda_{C_{1}} and λC2\lambda_{C_{2}} be the homology classes of their oriented longitudes. We see that λC1=4μK\lambda_{C_{1}}=4\mu_{K} and λC2=3μK\lambda_{C_{2}}=3\mu_{K}.

Let μK1\mu_{K_{1}}, μC1n\mu_{C_{1}^{n}} and μC2n\mu_{C_{2}^{n}} be the homology classes of meridians of K1K_{1}, C1nC_{1}^{n} and C2nC_{2}^{n}. Then we have that μK1=μK\mu_{K_{1}}=\mu_{K}, μC1n=μC1+nλC1\mu_{C_{1}^{n}}=-\mu_{C_{1}}+n\lambda_{C_{1}}, μC2n=μC2+2λC2\mu_{C_{2}^{n}}=-\mu_{C_{2}}+2\lambda_{C_{2}}. Hence

μK=μK1,μC1=μC1n+4nμK1,μC2=μC2n+6μK1.\mu_{K}=\mu_{K_{1}},\quad\mu_{C_{1}}=-\mu_{C_{1}^{n}}+4n\mu_{K_{1}},\quad\mu_{C_{2}}=-\mu_{C_{2}^{n}}+6\mu_{K_{1}}.

Thus we have the relation between the Alexander polynomials as

(3.1) ΔK1C1nC2n(x,y,z)=ΔL(x,x4ny1,x6z1).\Delta_{K_{1}\cup C_{1}^{n}\cup C_{2}^{n}}(x,y,z)=\Delta_{L}(x,x^{4n}y^{-1},x^{6}z^{-1}).

Since lk(K1,C2n)=lk(K,C2)=3\mathrm{lk}(K_{1},C_{2}^{n})=\mathrm{lk}(K,C_{2})=3 and lk(C1n,C2n)=lk(C1,C2)=0\mathrm{lk}(C_{1}^{n},C_{2}^{n})=\mathrm{lk}(C_{1},C_{2})=0, the Torres condition [36] gives

ΔK1C1nC2n(x,y,1)\displaystyle\Delta_{K_{1}\cup C_{1}^{n}\cup C_{2}^{n}}(x,y,1) =(x3y01)ΔK1C1n(x,y)\displaystyle=(x^{3}y^{0}-1)\Delta_{K_{1}\cup C_{1}^{n}}(x,y)
=(x31)ΔK1C1n(x,y).\displaystyle=(x^{3}-1)\Delta_{K_{1}\cup C_{1}^{n}}(x,y).

Furthermore, since lk(K1,C1n)=lk(K,C1)=4\mathrm{lk}(K_{1},C_{1}^{n})=\mathrm{lk}(K,C_{1})=4,

ΔK1C1n(x,1)=x41x1ΔK1(x).\Delta_{K_{1}\cup C_{1}^{n}}(x,1)=\frac{x^{4}-1}{x-1}\Delta_{K_{1}}(x).

Thus

ΔK1(x)=x1x41ΔK1C1n(x,1)=x1(x41)(x31)ΔK1C1nC2n(x,1,1).\Delta_{K_{1}}(x)=\frac{x-1}{x^{4}-1}\Delta_{K_{1}\cup C_{1}^{n}}(x,1)=\frac{x-1}{(x^{4}-1)(x^{3}-1)}\Delta_{K_{1}\cup C_{1}^{n}\cup C_{2}^{n}}(x,1,1).

Then the relation (3.1) gives

ΔK1(t)=t1(t41)(t31)ΔL(t,t4n,t6)=t1(t41)(t31)(t12n+18t12n+15+t8n+15t8n+14+t8n+11t8n+9+t4n+9t4n+7+t4n+4t4n+3+t31)=1t3+t2+t+11t31(t12n+15(t31)+t8n+9(t61)t8n+11(t31)+t4n+3(t61)t4n+4(t31)+(t31))=1t3+t2+t+1(t12n+15t8n+11t4n+4+1+(t8n+9+t4n+3)(t3+1)).\begin{split}\Delta_{K_{1}}(t)&=\frac{t-1}{(t^{4}-1)(t^{3}-1)}\Delta_{L}(t,t^{4n},t^{6})\\ &=\frac{t-1}{(t^{4}-1)(t^{3}-1)}(t^{12n+18}-t^{12n+15}+t^{8n+15}-t^{8n+14}+t^{8n+11}-t^{8n+9}\\ &\quad+t^{4n+9}-t^{4n+7}+t^{4n+4}-t^{4n+3}+t^{3}-1)\\ &=\frac{1}{t^{3}+t^{2}+t+1}\cdot\frac{1}{t^{3}-1}(t^{12n+15}(t^{3}-1)+t^{8n+9}(t^{6}-1)-t^{8n+11}(t^{3}-1)\\ &\quad+t^{4n+3}(t^{6}-1)-t^{4n+4}(t^{3}-1)+(t^{3}-1))\\ &=\frac{1}{t^{3}+t^{2}+t+1}(t^{12n+15}-t^{8n+11}-t^{4n+4}+1+(t^{8n+9}+t^{4n+3})(t^{3}+1)).\end{split}

We put

A1\displaystyle A_{1} =i=0n(t8n+12+4it8n+11+4i),\displaystyle=\sum_{i=0}^{n}(t^{8n+12+4i}-t^{8n+11+4i}),
A2\displaystyle A_{2} =t8n+9t8n+8,\displaystyle=t^{8n+9}-t^{8n+8},
A3\displaystyle A_{3} =i=0n(t4n+6+4it4n+4+4i),\displaystyle=\sum_{i=0}^{n}(t^{4n+6+4i}-t^{4n+4+4i}),
A4\displaystyle A_{4} =t4n+3t4n+1,\displaystyle=t^{4n+3}-t^{4n+1},
A5\displaystyle A_{5} =i=0n1(t4+4it1+4i).\displaystyle=\sum_{i=0}^{n-1}(t^{4+4i}-t^{1+4i}).

Then a direct calculation shows

(t3+t2+t+1)A1\displaystyle(t^{3}+t^{2}+t+1)A_{1} =t12n+15t8n+11,\displaystyle=t^{12n+15}-t^{8n+11},
(t3+t2+t+1)A2\displaystyle(t^{3}+t^{2}+t+1)A_{2} =t8n+12t8n+8,\displaystyle=t^{8n+12}-t^{8n+8},
(t3+t2+t+1)A3\displaystyle(t^{3}+t^{2}+t+1)A_{3} =(t8n+9t4n+5)+(t8n+8t4n+4),\displaystyle=(t^{8n+9}-t^{4n+5})+(t^{8n+8}-t^{4n+4}),
(t3+t2+t+1)A4\displaystyle(t^{3}+t^{2}+t+1)A_{4} =(t4n+6t4n+2)+(t4n+5t4n+1),\displaystyle=(t^{4n+6}-t^{4n+2})+(t^{4n+5}-t^{4n+1}),
(t3+t2+t+1)A5\displaystyle(t^{3}+t^{2}+t+1)A_{5} =(t4n+3t3)+(t4n+2t2)+(t4n+1t).\displaystyle=(t^{4n+3}-t^{3})+(t^{4n+2}-t^{2})+(t^{4n+1}-t).

Thus

(t3+t2+t+1)(A1+A2+A3+A4+A5+1)=t12n+15t8n+11t4n+4+1+t8n+12+t8n+9+t4n+6+t4n+3.\begin{split}(t^{3}+t^{2}+t+1)(A_{1}+A_{2}+A_{3}+A_{4}&+A_{5}+1)=t^{12n+15}-t^{8n+11}-t^{4n+4}+1\\ &+t^{8n+12}+t^{8n+9}+t^{4n+6}+t^{4n+3}.\end{split}

We have the conclusion ΔK1(t)=A1+A2+A3+A4+A5+1\Delta_{K_{1}}(t)=A_{1}+A_{2}+A_{3}+A_{4}+A_{5}+1 as desired. ∎

Theorem 3.2.

The Alexander polynomial of K2K_{2} is given as

ΔK2(t)=i=0n(t8n+12+4it8n+11+4i)+(t8n+9t8n+8)+i=02n1(t4n+8+2it4n+7+2i)+(t4n+6t4n+4)+(t4n+3t4n+1)+i=0n1(t4+4it1+4i)+1.\begin{split}\Delta_{K_{2}}(t)&=\sum_{i=0}^{n}(t^{8n+12+4i}-t^{8n+11+4i})+(t^{8n+9}-t^{8n+8})+\sum_{i=0}^{2n-1}(t^{4n+8+2i}-t^{4n+7+2i})\\ &\quad+(t^{4n+6}-t^{4n+4})+(t^{4n+3}-t^{4n+1})+\sum_{i=0}^{n-1}(t^{4+4i}-t^{1+4i})+1.\end{split}
Proof.

The argument is very similar to the proof of Theorem 3.1, so we omit the details.

Let L=KC1C2L=K\cup C_{1}\cup C_{2} be the oriented link illustrated in Figure 1(2). Its multivariable Alexander polynomial is

ΔL(x,y,z)=x6y3z2x3y3z2+x4y2z+x5yz+x3y2z2x3y2zx4yzx2y2z2+x4y+x2y2z+x3yzx3yxy2zx2yz+x31.\begin{split}\Delta_{L}(x,y,z)&=x^{6}y^{3}z^{2}-x^{3}y^{3}z^{2}+x^{4}y^{2}z+x^{5}yz+x^{3}y^{2}z^{2}-x^{3}y^{2}z\\ &-x^{4}yz-x^{2}y^{2}z^{2}+x^{4}y+x^{2}y^{2}z+x^{3}yz-x^{3}y-xy^{2}z-x^{2}yz+x^{3}-1.\end{split}

where x,y,zx,y,z correspond to the meridians of K,C1,C2K,C_{1},C_{2}, respectively.

Then

ΔK2(t)=t1(t41)(t31)ΔL(t,t4n,t6)=t1(t41)(t31)(t12n+18t12n+15+t8n+10+t4n+11+t8n+15t8n+9t4n+10t8n+14+t4n+4+t8n+8+t4n+9t4n+3t8n+7t4n+8+t31)=1t3+t2+t+11t31((t12n+15+t8n+7+t4n+8+1)(t31)+(t8n+9t4n+4t8n+8+t4n+3)(t61))=1t3+t2+t+1(t12n+15+t8n+7+t4n+8+1+(t8n+9t4n+4t8n+8+t4n+3)(t3+1)).\begin{split}\Delta_{K_{2}}(t)&=\frac{t-1}{(t^{4}-1)(t^{3}-1)}\Delta_{L}(t,t^{4n},t^{6})\\ &=\frac{t-1}{(t^{4}-1)(t^{3}-1)}(t^{12n+18}-t^{12n+15}+t^{8n+10}+t^{4n+11}+t^{8n+15}\\ &\quad-t^{8n+9}-t^{4n+10}-t^{8n+14}+t^{4n+4}+t^{8n+8}+t^{4n+9}-t^{4n+3}\\ &\quad-t^{8n+7}-t^{4n+8}+t^{3}-1)\\ &=\frac{1}{t^{3}+t^{2}+t+1}\cdot\frac{1}{t^{3}-1}((t^{12n+15}+t^{8n+7}+t^{4n+8}+1)(t^{3}-1)\\ &\quad+(t^{8n+9}-t^{4n+4}-t^{8n+8}+t^{4n+3})(t^{6}-1))\\ &=\frac{1}{t^{3}+t^{2}+t+1}(t^{12n+15}+t^{8n+7}+t^{4n+8}+1\\ &\qquad+(t^{8n+9}-t^{4n+4}-t^{8n+8}+t^{4n+3})(t^{3}+1)).\end{split}

Again, we put

B1\displaystyle B_{1} =i=0n(t8n+12+4it8n+11+4i),\displaystyle=\sum_{i=0}^{n}(t^{8n+12+4i}-t^{8n+11+4i}),
B2\displaystyle B_{2} =t8n+9t8n+8,\displaystyle=t^{8n+9}-t^{8n+8},
B3\displaystyle B_{3} =i=02n1(t4n+8+2it4n+7+2i),\displaystyle=\sum_{i=0}^{2n-1}(t^{4n+8+2i}-t^{4n+7+2i}),
B4\displaystyle B_{4} =t4n+6t4n+4,\displaystyle=t^{4n+6}-t^{4n+4},
B5\displaystyle B_{5} =t4n+3t4n+1,\displaystyle=t^{4n+3}-t^{4n+1},
B6\displaystyle B_{6} =i=0n1(t4+4it1+4i).\displaystyle=\sum_{i=0}^{n-1}(t^{4+4i}-t^{1+4i}).

A direct calculation shows

(t3+t2+t+1)B1\displaystyle(t^{3}+t^{2}+t+1)B_{1} =t12n+15t8n+11,\displaystyle=t^{12n+15}-t^{8n+11},
(t3+t2+t+1)B2\displaystyle(t^{3}+t^{2}+t+1)B_{2} =t8n+12t8n+8,\displaystyle=t^{8n+12}-t^{8n+8},
(t3+t2+t+1)B3\displaystyle(t^{3}+t^{2}+t+1)B_{3} =(t8n+9t4n+9)+(t8n+7t4n+7),\displaystyle=(t^{8n+9}-t^{4n+9})+(t^{8n+7}-t^{4n+7}),
(t3+t2+t+1)B4\displaystyle(t^{3}+t^{2}+t+1)B_{4} =(t4n+9t4n+5)+(t4n+8t4n+4),\displaystyle=(t^{4n+9}-t^{4n+5})+(t^{4n+8}-t^{4n+4}),
(t3+t2+t+1)B5\displaystyle(t^{3}+t^{2}+t+1)B_{5} =(t4n+6t4n+2)+(t4n+5t4n+1),\displaystyle=(t^{4n+6}-t^{4n+2})+(t^{4n+5}-t^{4n+1}),
(t3+t2+t+1)B6\displaystyle(t^{3}+t^{2}+t+1)B_{6} =(t4n+3t3)+(t4n+2t2)+(t4n+1t).\displaystyle=(t^{4n+3}-t^{3})+(t^{4n+2}-t^{2})+(t^{4n+1}-t).

This shows that

(t3+t2+t+1)(B1+B2+B3+B4+B5+B6+1)=t12n+15+t8n+7+t4n+8+1+(t8n+9t4n+4t8n+8+t4n+3)(t3+1).\begin{split}(t^{3}+t^{2}+t+1)&(B_{1}+B_{2}+B_{3}+B_{4}+B_{5}+B_{6}+1)=t^{12n+15}+t^{8n+7}+t^{4n+8}+1\\ &\quad+(t^{8n+9}-t^{4n+4}-t^{8n+8}+t^{4n+3})(t^{3}+1).\end{split}

Thus ΔK2(t)=B1+B2+B3+B4+B5+B6+1\Delta_{K_{2}}(t)=B_{1}+B_{2}+B_{3}+B_{4}+B_{5}+B_{6}+1 as desired. ∎

We recall the notion of formal semigroup for an L–space knot [37]. Let KK be an L–space knot in the 33–sphere. Then the Alexander polynomial of KK has a form of

(3.2) ΔK(t)=1ta1+ta2+tak1+tak,\Delta_{K}(t)=1-t^{a_{1}}+t^{a_{2}}+\dots-t^{a_{k-1}}+t^{a_{k}},

where 1=a1<a2<<ak=2g(K)1=a_{1}<a_{2}<\dots<a_{k}=2g(K), and g(K)g(K) is the genus of KK [27]. We expand the Alexander function into a formal power series as

(3.3) ΔK(t)1t=s𝒮Kts.\frac{\Delta_{K}(t)}{1-t}=\sum_{s\in\mathcal{S}_{K}}t^{s}.

(This is called the Milnor torsion in [10].) The set 𝒮K\mathcal{S}_{K} is a subset of non-negative integers, called the formal semigroup of KK. For example, for a torus knot T(p,q)(1<p<q)T(p,q)\ (1<p<q), its formal semigroup is known to be the actual semigroup of rank two,

p,q={ap+bqa,b0}\langle p,q\rangle=\{ap+bq\mid a,b\geq 0\}

(see [7, 37]). If an L–space knot is an iterated torus knot, then its formal semigroup is also a semigroup [37], but in general, the formal semigroup of a hyperbolic L–space knot is hardly a semigroup [3, 35].

Let m={iim}\mathbb{Z}_{\geq m}=\{i\in\mathbb{Z}\mid i\geq m\} and <0={ii<0}\mathbb{Z}_{<0}=\{i\in\mathbb{Z}\mid i<0\}.

Proposition 3.3.

The formal semigroup of K1K_{1} is given as

𝒮K1={0,4,8,,4n}{4n+3}{4n+6,4n+7,4n+10,4n+11,,8n+6,8n+7}{8n+9,8n+10}{8n+12,8n+13,8n+14,8n+16,8n+17,8n+18,,12n+8,12n+9,12n+10}12n+12.\begin{split}\mathcal{S}_{K_{1}}&=\{0,4,8,\dots,4n\}\cup\{4n+3\}\\ &\quad\cup\{4n+6,4n+7,4n+10,4n+11,\dots,8n+6,8n+7\}\cup\{8n+9,8n+10\}\\ &\quad\cup\{8n+12,8n+13,8n+14,8n+16,8n+17,8n+18,\\ &\qquad\dots,12n+8,12n+9,12n+10\}\cup\mathbb{Z}_{\geq 12n+12}.\end{split}
Proof.

We use A1,A2,,A5A_{1},A_{2},\dots,A_{5} in the proof of Theorem 3.1. For

ΔK11t=A11t+A21t+A31t+A41t+A51t+11t,\frac{\Delta_{K_{1}}}{1-t}=\frac{A_{1}}{1-t}+\frac{A_{2}}{1-t}+\frac{A_{3}}{1-t}+\frac{A_{4}}{1-t}+\frac{A_{5}}{1-t}+\frac{1}{1-t},

we expand each term as follows;

A11t\displaystyle\frac{A_{1}}{1-t} =i=0nt8n+11+4i,\displaystyle=-\sum_{i=0}^{n}t^{8n+11+4i},
A21t\displaystyle\frac{A_{2}}{1-t} =t8n+8,\displaystyle=-t^{8n+8},
A31t\displaystyle\frac{A_{3}}{1-t} =i=0n(t4n+5+4i+t4n+4+4i),\displaystyle=-\sum_{i=0}^{n}(t^{4n+5+4i}+t^{4n+4+4i}),
A41t\displaystyle\frac{A_{4}}{1-t} =t4n+2t4n+1,\displaystyle=-t^{4n+2}-t^{4n+1},
A51t\displaystyle\frac{A_{5}}{1-t} =i=0n1(t3+4i+t2+4i+t1+4i),\displaystyle=-\sum_{i=0}^{n-1}(t^{3+4i}+t^{2+4i}+t^{1+4i}),
11t\displaystyle\frac{1}{1-t} =1+t+t2+t3+.\displaystyle=1+t+t^{2}+t^{3}+\dots.

The conclusion immediately follows from these. ∎

Proposition 3.4.

The formal semigroup of K2K_{2} is given as

𝒮K2={0,4,8,,4n}{4n+3}{4n+6,4n+8,4n+10,,8n+4}{8n+6,8n+7,8n+9,8n+10}{8n+12,8n+13,8n+14,8n+16,8n+17,8n+18,,12n+8,12n+9,12n+10}12n+12.\begin{split}\mathcal{S}_{K_{2}}&=\{0,4,8,\dots,4n\}\cup\{4n+3\}\\ &\quad\cup\{4n+6,4n+8,4n+10,\dots,8n+4\}\cup\{8n+6,8n+7,8n+9,8n+10\}\\ &\quad\cup\{8n+12,8n+13,8n+14,8n+16,8n+17,8n+18,\\ &\qquad\dots,12n+8,12n+9,12n+10\}\cup\mathbb{Z}_{\geq 12n+12}.\end{split}
Proof.

The argument is similar to the proof of Proposition 3.3. For B1,B2,,B6B_{1},B_{2},\dots,B_{6} in the proof of Theorem 3.2, we expand

B11t\displaystyle\frac{B_{1}}{1-t} =i=0nt8n+11+4i,\displaystyle=-\sum_{i=0}^{n}t^{8n+11+4i},
B21t\displaystyle\frac{B_{2}}{1-t} =t8n+8,\displaystyle=-t^{8n+8},
B31t\displaystyle\frac{B_{3}}{1-t} =i=02n1t4n+7+2i,\displaystyle=-\sum_{i=0}^{2n-1}t^{4n+7+2i},
B41t\displaystyle\frac{B_{4}}{1-t} =t4n+4t4n+5,\displaystyle=-t^{4n+4}-t^{4n+5},
B51t\displaystyle\frac{B_{5}}{1-t} =t4n+1t4n+2,\displaystyle=-t^{4n+1}-t^{4n+2},
B61t\displaystyle\frac{B_{6}}{1-t} =i=0n1(t3+4i+t2+4i+t1+4i).\displaystyle=-\sum_{i=0}^{n-1}(t^{3+4i}+t^{2+4i}+t^{1+4i}).

Then the conclusion follows from these again. ∎

Corollary 3.5.

For i=1,2i=1,2, the formal semigroup of KiK_{i} is not a semigroup.

Proof.

By Propositions 3.3 and 3.4, we see that 4𝒮Ki4\in\mathcal{S}_{K_{i}} but 4n+4𝒮Ki4n+4\not\in\mathcal{S}_{K_{i}}. Hence 𝒮Ki\mathcal{S}_{K_{i}} is not closed under the addition, so is not a semigroup. ∎

Lemma 3.6.

Both of K1K_{1} and K2K_{2} are hyperbolic.

Proof.

By Corollary 3.5, the formal semigroup of KiK_{i} is not a semigroup. Hence KiK_{i} is not a torus knot, because the formal semigroup of a torus knot is a semigroup.

Assume for a contradiction that KiK_{i} is a satellite knot. Since KiK_{i} is the closure of a 44–braid, its bridge number is at most four. By [32], it is equal to four. Moreover, the companion is a 22–bridge knot and the pattern knot has wrapping number two. We know that both of the companion and the pattern knot are L–space knots and the pattern is braided by [5, 16]. Thus the companion is a 22–bridge torus knot [27], and KiK_{i} is its 22–cable. By [37], the formal semigroup of an iterated torus L–space knot is a semigroup, which contradicts Corollary 3.5. We have thus shown that KiK_{i} is hyperbolic. ∎

4. Upsilon invariants

In this section, we verify that the Upsilon invariants of K1K_{1} and K2K_{2} are the same. We will not calculate the Upsilon invariants. Instead, we determine the gap functions defined later. For an L–space knot, the Upsilon invariant is the Legendre–Fenchel transform of the gap function [6]. Hence if the gap functions of K1K_{1} and K2K_{2} share the same convex hull, then their Upsilon invariants also coincide.

First, we quickly review the Legendre–Fenchel transformation.

For a function f:f\colon\mathbb{R}\to\mathbb{R}, the Legendre–Fenchel transform f:{}f^{*}\colon\mathbb{R}\to\mathbb{R}\cup\{\infty\} is defined as

f(t)=supx{txf(x)}.f^{*}(t)=\sup_{x\in\mathbb{R}}\{tx-f(x)\}.

The domain of ff^{*} is the set {tf(t)<}\{t\mid f^{*}(t)<\infty\}.

The Legendre transform is defined only for differentiable convex functions, but the Legendre–Fenchel transform can be defined even for non-convex functions with non-differentiable points. The transform ff^{*} is always a convex function. Hence, if ff is not convex, then the double Legendre–Fenchel transform ff^{**} does not return ff. In this case, ff^{**} gives the convex hull of the function ff. Thus we see that ff^{*} depends only on the convex hull of ff.

Next, we recall the notion of gap function introduced in [7].

Let KK be an L–space knot with formal semigroup 𝒮K\mathcal{S}_{K}. Then 𝒢K=𝒮K\mathcal{G}_{K}=\mathbb{Z}-\mathcal{S}_{K} is called the gap set. In fact, 𝒢K=<0{a1,a2,,ag}\mathcal{G}_{K}=\mathbb{Z}_{<0}\cup\{a_{1},a_{2},\dots,a_{g}\}, where g=g(K)g=g(K), and 0<a1<a2<<ag0<a_{1}<a_{2}<\dots<a_{g}. The part a1,a2,,aga_{1},a_{2},\dots,a_{g} is called the gap sequence. Then it is easy to restore the Alexander polynomial as

ΔK(t)=1+(t1)(ta1+ta2++tag).\Delta_{K}(t)=1+(t-1)(t^{a_{1}}+t^{a_{2}}+\dots+t^{a_{g}}).

From the gap set 𝒢K\mathcal{G}_{K}, we define the function I:0I\colon\mathbb{Z}\to\mathbb{Z}_{\geq 0} by

I(m)=#{i𝒢Kim},I(m)=\#\{i\in\mathcal{G}_{K}\mid i\geq m\},

and let J(m)=I(m+g)J(m)=I(m+g). Then we extend J(m)J(m) linearly to obtain a piecewise linear function on \mathbb{R}. That is, for kk\in\mathbb{Z}, if J(k)=J(k+1)J(k)=J(k+1), then J(x)=J(k)J(x)=J(k) on [k,k+1][k,k+1], and if J(k+1)=J(k)1J(k+1)=J(k)-1, then J(k+x)=J(k)xJ(k+x)=J(k)-x for 0x10\leq x\leq 1. Borodzik and Hedden [6] showed that the Upsilon invariant of KK is the Legendre–Fenchel transform of the function 2J(m)2J(-m). We call this function 2J(m)2J(-m) the gap function of KK.

Example 4.1.

Let KK be the (2,3,7)(-2,3,7)–pretzel knot. It admits a lens space surgery, so is an L–space knot. Also, it has genus 55. The Alexander polynomial ΔK(t)\Delta_{K}(t) is 1t+t3t4+t5t6+t7t9+t101-t+t^{3}-t^{4}+t^{5}-t^{6}+t^{7}-t^{9}+t^{10}. Then 𝒮K={0,3,5,7,8}10\mathcal{S}_{K}=\{0,3,5,7,8\}\cup\mathbb{Z}_{\geq 10}, and 𝒢K=<0{1,2,4,6,9}\mathcal{G}_{K}=\mathbb{Z}_{<0}\cup\{1,2,4,6,9\}. Tables 1 and 2 show the values of I(m)I(m) and the gap function 2J(m)2J(-m).

mm 10\geq 10 9 8 7 6 5 4 3 2 1 0 1-1 2-2 \dots
I(m)I(m) 0 1 1 1 2 2 3 3 4 5 5 6 7 \dots
Table 1. I(m)I(m) for the (2,3,7)(-2,3,7)–pretzel knot.
mm 5\leq-5 4-4 3-3 2-2 1-1 0 1 2 3 4 5 6 7 \dots
2J(m)2J(-m) 0 2 2 2 4 4 6 6 8 10 10 12 14 \dots
Table 2. The gap function 2J(m)2J(-m) for the (2,3,7)(-2,3,7)–pretzel knot.

Figure 2 shows the graph of the gap function 2J(m)2J(-m) and its convex hull (broken line). Here, the convex hull f(x)f(x) of the gap function is given by

f(x)={0for x5,23(x+5)for 5x2,x+4for 2x2,43(x5)+10for 2x5,2xfor 5x.f(x)=\begin{cases}0&\text{for $x\leq-5$},\\ \frac{2}{3}(x+5)&\text{for $-5\leq x\leq-2$},\\ x+4&\text{for $-2\leq x\leq 2$},\\ \frac{4}{3}(x-5)+10&\text{for $2\leq x\leq 5$},\\ 2x&\text{for $5\leq x$}.\end{cases}
Refer to caption
Figure 2. The graph of the gap function 2J(m)2J(-m) for the (2,3,7)(-2,3,7)–pretzel knot and its convex hull (broken line).

Then the Legendre–Fenchel transformation gives the Upsilon invariant

ΥK(t)={5tfor 0t23,2t2for 23t1,2t6for 1t43,5t10for 43t2.\Upsilon_{K}(t)=\begin{cases}-5t&\text{for $0\leq t\leq\frac{2}{3}$},\\ -2t-2&\text{for $\frac{2}{3}\leq t\leq 1$},\\ 2t-6&\text{for $1\leq t\leq\frac{4}{3}$},\\ 5t-10&\text{for $\frac{4}{3}\leq t\leq 2$}.\end{cases}

In general, the gap function of an L–space knot has a specific property.

  • The slope of each segment of the graph is 0 or 22.

Although this observation is easy to see, we will use it essentially in Section 6 with further investigation.

Now, we calculate the gap functions of K1K_{1} and K2K_{2}.

From Proposition 3.3, the gap set 𝒢K1\mathcal{G}_{K_{1}} is

<0{1,2,3,5,6,7,,4n3,4n2,4n1}{4n+1,4n+2}{4n+4,4n+5,4n+8,4n+9,4n+12,4n+13,,8n+4,8n+5}{8n+8}{8n+11,8n+15,8n+19,,12n+7,12n+11}.\begin{split}\mathbb{Z}_{<0}&\cup\{1,2,3,5,6,7,\dots,4n-3,4n-2,4n-1\}\cup\{4n+1,4n+2\}\\ &\cup\{4n+4,4n+5,4n+8,4n+9,4n+12,4n+13,\dots,8n+4,8n+5\}\\ &\cup\{8n+8\}\cup\{8n+11,8n+15,8n+19,\dots,12n+7,12n+11\}.\end{split}

Hence the values of I(m)I(m) is given as in Table 3. When mm is an integer not in the table, I(m)I(m) takes the same value as the nearest mm^{\prime} with m>mm^{\prime}>m. For example, I(m)=I(12n+11)=1I(m)=I(12n+11)=1 for m=12n+10,12n+9,12n+8m=12n+10,12n+9,12n+8.

mm 12n+12\geq 12n+12 12n+1112n+11 12n+712n+7 \dots 8n+118n+11 8n+88n+8
I(m)I(m) 0 1 2 \dots n+1n+1 n+2n+2
8n+58n+5 8n+48n+4 8n+18n+1 8n8n \dots 4n+54n+5 4n+44n+4 4n+24n+2 4n+14n+1
n+3n+3 n+4n+4 n+5n+5 n+6n+6 \dots 3n+33n+3 3n+43n+4 3n+53n+5 3n+63n+6
4n14n-1 4n24n-2 4n34n-3 \dots 3 2 1 1-1 2-2 \dots
3n+73n+7 3n+83n+8 3n+93n+9 \dots 6n+46n+4 6n+56n+5 6n+66n+6 6n+76n+7 6n+86n+8 \dots
Table 3. The function I(m)I(m) for K1K_{1}.

Let J(m)=I(m+g)J(m)=I(m+g) with g=6n+6g=6n+6. Then the gap function 2J(m)2J(-m) takes the values as in Table 4.

mm 6n6\leq-6n-6 6n5-6n-5 6n1-6n-1 \dots 2n5-2n-5 2n2-2n-2
2J(m)2J(-m) 0 2 4 \dots 2n+22n+2 2n+42n+4
2n+1-2n+1 2n+2-2n+2 2n+5-2n+5 2n+6-2n+6 \dots 2n+12n+1 2n+22n+2 2n+42n+4 2n+52n+5
2n+62n+6 2n+82n+8 2n+102n+10 2n+122n+12 \dots 6n+66n+6 6n+86n+8 6n+106n+10 6n+126n+12
2n+72n+7 2n+82n+8 2n+92n+9 \dots 6n+36n+3 6n+46n+4 6n+56n+5 6n+76n+7 6n+86n+8 \dots
6n+146n+14 6n+166n+16 6n+186n+18 \dots 12n+812n+8 12n+1012n+10 12n+1212n+12 12n+1412n+14 12n+1612n+16 \dots
Table 4. The gap function 2J(m)2J(-m) for K1K_{1}.

Figure 3 shows the graph of the gap function 2J(m)2J(-m) of K1K_{1} when n=1n=1.

Refer to caption
Figure 3. The graph of the gap function of K1K_{1} with n=1n=1.

Similarly, the gap set 𝒢K2\mathcal{G}_{K_{2}} is

<0{1,2,3,5,6,7,,4n3,4n2,4n1}{4n+1,4n+2}{4n+4,4n+5}{4n+7,4n+9,,8n+3,8n+5}{8n+8,8n+11}{8n+15,8n+19,,12n+7,12n+11}\begin{split}\mathbb{Z}_{<0}&\cup\{1,2,3,5,6,7,\dots,4n-3,4n-2,4n-1\}\cup\{4n+1,4n+2\}\\ &\cup\{4n+4,4n+5\}\cup\{4n+7,4n+9,\dots,8n+3,8n+5\}\\ &\cup\{8n+8,8n+11\}\cup\{8n+15,8n+19,\dots,12n+7,12n+11\}\end{split}

from Proposition 3.4.

The values of I(m)I(m) and the gap function 2J(m)2J(-m) are given as in Tables 5 and 6.

mm 12n+12\geq 12n+12 12n+1112n+11 12n+712n+7 \dots 8n+158n+15 8n+118n+11 8n+88n+8
I(m)I(m) 0 1 2 \dots nn n+1n+1 n+2n+2
8n+58n+5 8n+38n+3 8n+18n+1 \dots 4n+74n+7 4n+54n+5 4n+44n+4 4n+24n+2 4n+14n+1
n+3n+3 n+4n+4 n+5n+5 \dots 3n+23n+2 3n+33n+3 3n+43n+4 3n+53n+5 3n+63n+6
4n14n-1 4n24n-2 4n34n-3 \dots 3 2 1 1-1 2-2 \dots
3n+73n+7 3n+83n+8 3n+93n+9 \dots 6n+46n+4 6n+56n+5 6n+66n+6 6n+76n+7 6n+86n+8 \dots
Table 5. The function I(m)I(m) for K2K_{2}.
mm 6n6\leq-6n-6 6n5-6n-5 6n1-6n-1 \dots 2n9-2n-9 2n5-2n-5 2n2-2n-2
2J(m)2J(-m) 0 2 4 \dots 2n2n 2n+22n+2 2n+42n+4
2n+1-2n+1 2n+3-2n+3 2n+5-2n+5 \dots 2n12n-1 2n+12n+1 2n+22n+2 2n+42n+4 2n+52n+5
2n+62n+6 2n+82n+8 2n+102n+10 \dots 6n+46n+4 6n+66n+6 6n+86n+8 6n+106n+10 6n+126n+12
2n+72n+7 2n+82n+8 2n+92n+9 \dots 6n+36n+3 6n+46n+4 6n+56n+5 6n+76n+7 6n+86n+8 \dots
6n+146n+14 6n+166n+16 6n+186n+18 \dots 12n+812n+8 12n+1012n+10 12n+1212n+12 12n+1412n+14 12n+1612n+16 \dots
Table 6. The gap function 2J(m)2J(-m) for K2K_{2}.
Refer to caption
Figure 4. The parts of the graph of a gap function. The broken lines show the parts of convex hull with slope ss.
Lemma 4.2.

For i=1,2i=1,2, the convex hull f(x)f(x) of the gap function 2J(m)2J(-m) for KiK_{i} is given by

f(x)={0for x6n6,12(x+6n+6)for 6n6x2n6,23(x+2n+6)+2nfor 2n6x2n,x+4n+4for 2nx2n,43(x2n)+6n+4for 2nx2n+6,32(x2n6)+6n+12for 2n+6x6n+6,2xfor 6n+6x.f(x)=\begin{cases}0&\text{for $x\leq-6n-6$},\\ \frac{1}{2}(x+6n+6)&\text{for $-6n-6\leq x\leq-2n-6$},\\ \frac{2}{3}(x+2n+6)+2n&\text{for $-2n-6\leq x\leq-2n$},\\ x+4n+4&\text{for $-2n\leq x\leq 2n$},\\ \frac{4}{3}(x-2n)+6n+4&\text{for $2n\leq x\leq 2n+6$},\\ \frac{3}{2}(x-2n-6)+6n+12&\text{for $2n+6\leq x\leq 6n+6$},\\ 2x&\text{for $6n+6\leq x$}.\end{cases}
Proof.

Consider the gap function of K1K_{1}. Let ff be the convex hull. From Table 4, it is obvious that f(x)=0f(x)=0 for x6n6x\leq-6n-6 and f(x)=2xf(x)=2x for x6n+6x\geq 6n+6.

On the interval [6n6,6n2][-6n-6,-6n-2], the gap function has the branch as shown in Figure 4(b). It repeats on the intervals [6n2,6n+2],,[2n10,2n6][-6n-2,-6n+2],\dots,[-2n-10,-2n-6]. Thus f(x)=12(x+6n+6)f(x)=\frac{1}{2}(x+6n+6) on [6n6,2n6][-6n-6,-2n-6].

On [2n6,2n3][-2n-6,-2n-3] and [2n3,2n][-2n-3,-2n], the branch is of Figure 4(c). Hence f(x)=23(x+2n+6)+2nf(x)=\frac{2}{3}(x+2n+6)+2n on [2n6,2n][-2n-6,-2n].

Similarly, the branch of Figure 4(d) repeats on the intervals [2n,2n+4],[2n+4,2n+6],,[2n4,2n][-2n,2n+4],[-2n+4,-2n+6],\dots,[2n-4,2n]. This gives f(x)=x+4n+4f(x)=x+4n+4 on [2n,2n][-2n,2n].

On [2n,2n+3][2n,2n+3] and [2n+3,2n+6][2n+3,2n+6], the branch of Figure 4(f) appears. Thus f(x)=43(x2n)+6n+4f(x)=\frac{4}{3}(x-2n)+6n+4 on [2n,2n+6][2n,2n+6].

Finally, the branch of Figure 4(g) repeats on [2n+6,2n+10],,[6n+2,6n+6][2n+6,2n+10],\dots,[6n+2,6n+6]. Then f(x)=32(x2n6)+6n+12f(x)=\frac{3}{2}(x-2n-6)+6n+12 on [2n+6,6n+6][2n+6,6n+6]. We have thus shown that the convex hull f(x)f(x) is given as claimed for K1K_{1}.

Next, consider the gap function of K2K_{2}. For x2nx\leq-2n, the situation is the same as K1K_{1}.

On [2n,2n+2][-2n,-2n+2], the branch of Figure 4(e) appears. This branch repeats on [2n+2,2n+4],,[2n2,2n][-2n+2,-2n+4],\dots,[2n-2,2n]. However, the convex hull is the same as K1K_{1}.

For the remaining range x2nx\geq 2n, the gap function is the same as one of K1K_{1}. In conclusion, the gap functions of K1K_{1} and K2K_{2} are distinct only on [2n,2n][-2n,2n], but their convex hulls coincide there. ∎

Corollary 4.3.

The Upsilon invariants of K1K_{1} and K2K_{2} coincide.

Proof.

The Upsilon invariant is the Legendre–Fenchel transform of the gap function 2J(m)2J(-m). In fact, it depends only on the convex hull of the gap function. By Lemma 4.2, K1K_{1} and K2K_{2} have the same convex hull for their gap functions. Thus the conclusion follows. ∎

5. The Montesinos trick

In this section, we verify that K1K_{1} and K2K_{2} admit positive Dehn surgeries yielding L–spaces by using the Montesinos trick [24]. For a surgery diagram on a strongly invertible link, the Montesinos trick describes the resulting closed 33–manifold as the double branched cover of another knot or link obtained from tangle replacements corresponding to the surgery coefficients on some link obtained from the quotient of the original strongly invertible link under the strong involution (see also [26, 41]).

In Figure 1(1) and (2), each link KC1C2K\cup C_{1}\cup C_{2} is placed in a strongly invertible position, where the dotted line indicates the axis of the involution.

Lemma 5.1.

For K1K_{1}, (16n+21)(16n+21)–surgery yields an L–space.

Proof.

Assign the surgery coefficient 33 on KK in Figure 1(1). After performing (1/n)(-1/n)–surgery on C1C_{1} and (1/2)(-1/2)–surgery on C2C_{2}, our knot K1K_{1} has surgery coefficient 16n+2116n+21.

The left of Figure 5 shows the knot obtained from the tangle replacements. In the diagram of Figure 1, we should remark that the component KK has writhe 33. Hence the tangle replacement corresponding to the quotient of KK is realized by the 0–tangle (depicted as the dotted circle).

Then Figures 5, 6, and 7 show the deformation of the knot. Finally, we obtain the Montesinos knot M(3/7,1/3,1/n)M(-3/7,-1/3,-1/n). Thus the double branched cover is the Seifert fibered manifold M=M(0;3/7,1/3,1/n)M=M(0;-3/7,-1/3,-1/n). We use the notation of [22]. That is, M(e0;r1,r2,r3)M(e_{0};r_{1},r_{2},r_{3}) is obtained by e0e_{0}–surgery on the unknot with three meridians having (/ri)(-/r_{i})–surgery on the ii-th one. Then M=M(0;3/7,1/3,1/n)-M=M(0;3/7,1/3,1/n). By the criterion of [21, 22], MM is an L–space. ∎

Refer to caption
Figure 5. The deformation for K1K_{1}. Each rectangle box contains horizontal right-handed half-twists with indicated number.
Refer to caption
Figure 6. The deformation for K1K_{1} (continued from Figure 5).
Refer to caption
Figure 7. The deformation for K1K_{1} (continued from Figure 6). The left bottom is the Montesinos knot M(3/7,1/3,1/n)M(-3/7,-1/3,-1/n).
Lemma 5.2.

For K2K_{2}, (16n+20)(16n+20)–surgery yields an L–space.

Proof.

Assign the surgery coefficient 22 on KK in Figure 1(2). After performing (1/n)(-1/n)–surgery on C1C_{1} and (1/2)(-1/2)–surgery on C2C_{2}, K2K_{2} has surgery coefficient 16n+2016n+20.

The process is similar to that for K1K_{1}. We should remark that the tangle replacement to the quotient of KK is realized by (1)(-1)–tangle as depicted in the dotted circle in Figure 8 (left), because KK has writhe 33 in the diagram.

Refer to caption
Figure 8. The deformation for K2K_{2}. Let \ell be the right link.

Let \ell be the link as illustrated in the right of Figure 8. We need to verify that the double branched cover of \ell is an L–space.

For the crossing of \ell encircled in Figure 8 (right), we perform two resolutions as shown in Figure 9. Let \ell_{\infty} and 0\ell_{0} be the resulting knots. It is straightforward to calculate det=16n+20\det\ell=16n+20, det=9\det\ell_{\infty}=9 and det0=16n+11\det\ell_{0}=16n+11 from the checkerboard colorings on the diagrams of Figures 8, 11 and 12. Thus the equation det=det+det0\det\ell=\det\ell_{\infty}+\det\ell_{0} holds. This implies that if the double branched covers of \ell_{\infty} and 0\ell_{0} are L–spaces, then so is the double branched cover of \ell ([8, 27, 28]).

Refer to caption
Figure 9. Two resolutions.
Claim 5.3.

The knot \ell_{\infty} is the (3,3,n1)(-3,3,n-1)–pretzel knot. Its double branched cover is an L–space.

Proof.

Figures 10 and 11 show that the knot \ell_{\infty} is the (3,3,n1)(-3,3,n-1)–pretzel knot.

If n=1n=1, then \ell_{\infty} is the connected sum of torus knots T(2,3)T(2,3) and T(2,3)T(2,-3). Then the double branched cover is the connected sum of lens spaces L(3,1)#L(3,1)L(3,1)\#L(3,-1), which is an L–space. If n=2n=2, then \ell_{\infty} is the 22–bridge knot S(4/9)S(4/9), so the double cover is a lens space. Hence we assume n>2n>2.

Since \ell_{\infty} is the Montesinos knot M(0;1/3,1/3,1/(n1))M(0;1/3,-1/3,-1/(n-1)), its double branched cover MM is the Seifert fibered manifold M(0;1/3,1/3,1/(n1))M(0;1/3,-1/3,-1/(n-1)). Then M=M(1;2/3,1/3,1/(n1))-M=M(-1;2/3,1/3,1/(n-1)).

We use the criterion of [22]. If n>3n>3, then set r1=2/3r_{1}=2/3, r2=1/3r_{2}=1/3 and r3=1/(n1)r_{3}=1/(n-1). Then 1r1r2r301\geq r_{1}\geq r_{2}\geq r_{3}\geq 0. If there are no coprime integers m>a>0m>a>0 such that a/m>r1a/m>r_{1}, (ma)/m>r2(m-a)/m>r_{2} and 1/m>r31/m>r_{3}, then M-M is an L–space. However, the first two give 2/3<a/m<2/32/3<a/m<2/3, so there are no such integers.

Finally, assume n=3n=3. Set r1=2/3r_{1}=2/3, r2=1/2r_{2}=1/2 and r3=1/3r_{3}=1/3. Then r1r2r3r_{1}\geq r_{2}\geq r_{3}. If 1/m>r3=1/31/m>r_{3}=1/3, then m<3m<3. For m=2m=2 and a=1a=1, a/m>r1=2/3a/m>r_{1}=2/3 does not hold. Thus there are no coprime integers m>a>0m>a>0 as desired, which implies that M-M is an L–space. ∎

Refer to caption
Figure 10. A deformation of the knot \ell_{\infty}.
Refer to caption
Figure 11. A deformation of the knot \ell_{\infty} (continued from Figure 10). The right is the (3,3,n1)(-3,3,n-1)–pretzel knot.
Claim 5.4.

The double branched cover of 0\ell_{0} is an L–space.

Proof.

For the crossing encircled in Figure 12, we further perform the resolutions, which yield 0\ell_{0\infty} and 00\ell_{00}. Clearly, 0=\ell_{0\infty}=\ell_{\infty}. Hence det0=9\mathrm{det}\,\ell_{0\infty}=9.

Refer to caption
Figure 12. The knot 0\ell_{0}.

We can confirm that 00\ell_{00} is the connected sum of the Hopf link and a Montesinos knot as shown in Figures 13 and 14. From the diagram of Figure 14, we see that det00=16n+2\mathrm{det}\,\ell_{00}=16n+2. Recall that det0=16n+11\mathrm{det}\,\ell_{0}=16n+11. Hence the equation det0=det0+det00\mathrm{det}\,\ell_{0}=\mathrm{det}\,\ell_{0\infty}+\mathrm{det}\,\ell_{00} holds.

Refer to caption
Figure 13. The knot 00\ell_{00} has the Hopf link as its connected summand.
Refer to caption
Figure 14. The knot 00\ell_{00} is the connected sum of the Hopf link and the Montesinos knot M(1/2,1/3,n/(2n+1))M(1/2,-1/3,n/(2n+1)).

From Claim 5.3, the double branched cover of 0\ell_{0\infty} is an L–space. It remains to show that the double branched cover of 00\ell_{00} is an L–space.

The double branched cover of the Montesinos knot M=M(1/2,1/3,n/(2n+1))M=M(1/2,-1/3,n/(2n+1)) is the Seifert fibered manifold M(0;1/2,1/3,n/(2n+1))M(0;1/2,-1/3,n/(2n+1)). Since MM is homeomorphic to M(1;1/2,2/3,n/(2n+1))M(-1;1/2,2/3,n/(2n+1)), set r1=2/3r_{1}=2/3, r2=1/2r_{2}=1/2 and r3=n/(2n+1)r_{3}=n/(2n+1). Then 1r1r2r301\geq r_{1}\geq r_{2}\geq r_{3}\geq 0. We apply the criterion of [22] again. If 1/m>r3=n/(2n+1)1/31/m>r_{3}=n/(2n+1)\geq 1/3, then m<3m<3. Hence there are no coprime integers m>a>0m>a>0 such that a/m>r1=2/3a/m>r_{1}=2/3. Thus MM is an L–space.

The double branched cover of 00\ell_{00} is the connected sum of a lens space L(2,1)L(2,1) and MM. Since the sum of L–spaces is an L–space [27], we have the conclusion. ∎

By Claims 5.3 and 5.4, we obtain that the double branched cover of \ell is an L–space. ∎

6. Restorability of Alexander polynomials

In this section, we investigate the restorability of Alexander polynomial of an L–space knot from the Upsilon invariant.

As easy examples, we examine two torus knots.

Example 6.1.

(1) Let K=T(3,4)K=T(3,4). Then ΔK(t)=1t+t3t5+t6\Delta_{K}(t)=1-t+t^{3}-t^{5}+t^{6}, so 𝒮K={0,3,4}6\mathcal{S}_{K}=\{0,3,4\}\cup\mathbb{Z}_{\geq 6} and 𝒢K=<0{1,2,5}\mathcal{G}_{K}=\mathbb{Z}_{<0}\cup\{1,2,5\}. It is easy to calculate ΥK(t)\Upsilon_{K}(t) as

ΥK(t)={3tfor 0t23,2for 23t43,3t6for 43t2.\Upsilon_{K}(t)=\begin{cases}-3t&\text{for $0\leq t\leq\frac{2}{3}$},\\ -2&\text{for $\frac{2}{3}\leq t\leq\frac{4}{3}$},\\ 3t-6&\text{for $\frac{4}{3}\leq t\leq 2$}.\end{cases}

The Legendre–Fenchel transformation on ΥK(t)\Upsilon_{K}(t) gives a function

f(x)={0for x3,23(x+3)for 3x0,43x+2for 0x3,2xfor 3x.f(x)=\begin{cases}0&\text{for $x\leq-3$},\\ \frac{2}{3}(x+3)&\text{for $-3\leq x\leq 0$},\\ \frac{4}{3}x+2&\text{for $0\leq x\leq 3$},\\ 2x&\text{for $3\leq x$}.\end{cases}

Of course, this is the convex hull of the gap function of KK. Figure 15 shows the graphs of gap function of KK and ff.

Refer to caption
Figure 15. The graphs of the gap functions and their convex hulls (broken line) of T(3,4)T(3,4) (left) and T(3,5)T(3,5) (right).

We consider the possibility of another gap function GG whose convex hull is ff. First, it forces G(3)=0G(-3)=0, G(0)=2G(0)=2 and G(3)=6G(3)=6. Recall that each segment of the graph of a gap function has slope 0 or 22 as mentioned in Section 4. Hence G(2)=2G(-2)=2. Since a gap function is increasing, G(1)=2G(-1)=2. Similarly, it is necessary that G(1)=4G(1)=4 and G(2)=6G(2)=6. Thus GG coincides with the gap function of KK.

This means that if another L–space knot KK^{\prime} has the same Upsilon invariant as KK, then ΔK(t)=ΔK(t)\Delta_{K^{\prime}}(t)=\Delta_{K}(t), because a gap function uniquely determines the Alexander polynomial.

(2) Let K=T(3,5)K=T(3,5). We have ΔK(t)=1t+t3t4+t5t7+t8\Delta_{K}(t)=1-t+t^{3}-t^{4}+t^{5}-t^{7}+t^{8}, so 𝒮K={0,3,5,6}8\mathcal{S}_{K}=\{0,3,5,6\}\cup\mathbb{Z}_{\geq 8}, and 𝒢K=<0{1,2,4,7}\mathcal{G}_{K}=\mathbb{Z}_{<0}\cup\{1,2,4,7\}. Then ΥK(t)\Upsilon_{K}(t) is given as

ΥK(t)={4tfor 0t23,t2for 23t1,t4for 1t43,4t8for 43t2.\Upsilon_{K}(t)=\begin{cases}-4t&\text{for $0\leq t\leq\frac{2}{3}$},\\ -t-2&\text{for $\frac{2}{3}\leq t\leq 1$},\\ t-4&\text{for $1\leq t\leq\frac{4}{3}$},\\ 4t-8&\text{for $\frac{4}{3}\leq t\leq 2$}.\end{cases}

Figure 15 shows the graphs of gap function of KK and the convex hull, which is the Legendre–Fenchel transform of ΥK(t)\Upsilon_{K}(t). As in (1), the convex hull uniquely restores the gap function.

In general, it is rare that the convex hull uniquely restores a gap function. In Example 4.1, we determined the gap function and its convex hull of the (2,3,7)(-2,3,7)–pretzel knot (see Figure 2). It is possible that another gap function GG takes the same values on integers except G(0)=6G(0)=6, keeping the same convex hull. This new gap function corresponds to the Alexander polynomial Δ(t)=1t+t3t5+t7t9+t10\Delta(t)=1-t+t^{3}-t^{5}+t^{7}-t^{9}+t^{10}. This polynomial satisfies the condition of [20], but there is no hyperbolic L–space knot in Dunfield’s list whose Alexander polynomial is Δ(t)\Delta(t). It seems to be a hard question whether there exists a hyperbolic L–space knot with Δ(t)\Delta(t). Of course, there exists a hyperbolic knot whose Alexander polynomial is Δ(t)\Delta(t) by [13, 34]. Also, Δ(t)\Delta(t) is the Alexander polynomial of the (2,3)(2,3)–cable of T(2,5)T(2,5), which is not an L–space knot [15].

If we put off the realizability of the Alexander polynomial or the gap function by a hyperbolic L–space knot, then we can easily design many Alexander polynomials which are restorable from convex hulls.

It is a classical result that any polynomial Δ(t)\Delta(t) satisfying Δ(1)=1\Delta(1)=1 and Δ(t1)=.Δ(t)\Delta(t^{-1})\stackrel{{\scriptstyle.}}{{=}}\Delta(t) is realized by a knot in the 33–sphere as its Alexander polynomial. (Here, =.\stackrel{{\scriptstyle.}}{{=}} shows the equality up to units ±ti\pm t^{i} in the Laurent polynomial ring [t,t1]\mathbb{Z}[t,t^{-1}].) Furthermore, we assume that Δ(t)\Delta(t) has the form of (3.2). Formally, we define the formal semigroup 𝒮\mathcal{S} by (3.3), and in turn, its gap set and the gap function.

Proposition 6.2.

Let m3m\geq 3 be an integer, and let Δ(t)=1t+tmtm+1+tm+2t2m+1+t2m+2\Delta(t)=1-t+t^{m}-t^{m+1}+t^{m+2}-t^{2m+1}+t^{2m+2}. Then its gap function, defined formally, is uniquely determined from the convex hull.

Again, the polynomial Δ(t)\Delta(t) in Proposition 6.2 satisfies the condition of [20], but it is open whether Δ(t)\Delta(t) is realized by a hyperbolic L–space knot or not. (When m=3m=3, Δ(t)\Delta(t) is the Alexander polynomial of T(3,5)T(3,5).)

Proof.

By (3.3), the formal semigroup is 𝒮={0,m}{m+2,m+3,,2m}2m+2\mathcal{S}=\{0,m\}\cup\{m+2,m+3,\dots,2m\}\cup\mathbb{Z}_{2m+2}, so the gap set is 𝒢=<0{1,2,,m1}{m+1,2m+1}\mathcal{G}=\mathbb{Z}_{<0}\cup\{1,2,\dots,m-1\}\cup\{m+1,2m+1\}. Set g=m+1g=m+1. Then we can calculate the gap function 2J(m)2J(-m) as in Table 7.

mm m1\leq-m-1 m-m 0 22 33 \dots mm m+2m+2 m+3m+3 \dots
2J(m)2J(-m) 0 2 4 66 88 \dots 2m+22m+2 2m+42m+4 2m+62m+6 \dots
Table 7. The gap function 2J(m)2J(-m).

Let ff be the convex hull. Then it is given by

f(x)={0for xm1,2m(x+m+1)for m1x1,x+3for 1x1,2m2m(x1)+4for 1xm+1,2xfor m+1x.f(x)=\begin{cases}0&\text{for $x\leq-m-1$},\\ \frac{2}{m}(x+m+1)&\text{for $-m-1\leq x\leq-1$},\\ x+3&\text{for $-1\leq x\leq 1$},\\ \frac{2m-2}{m}(x-1)+4&\text{for $1\leq x\leq m+1$},\\ 2x&\text{for $m+1\leq x$}.\end{cases}

Since each segment of the graph of any gap function has slope 0 or 22, there is no other gap function whose convex hull is ff. ∎

Finally, we prove Theorem 1.3. For reader’s convenience, we record the braid words for the knots t09847 and v2871. Both are the closures of 44–braids, whose words are almost the same:

(σ2σ1σ3σ2)3(σ2σ12σ2)σ1and(σ2σ1σ3σ2)3(σ2σ12σ2)σ13.(\sigma_{2}\sigma_{1}\sigma_{3}\sigma_{2})^{3}(\sigma_{2}\sigma_{1}^{2}\sigma_{2})\sigma_{1}\quad\text{and}\quad(\sigma_{2}\sigma_{1}\sigma_{3}\sigma_{2})^{3}(\sigma_{2}\sigma_{1}^{2}\sigma_{2})\sigma_{1}^{3}.
Proof of Theorem 1.3.

Let KK be the hyperbolic knot t09847 in the SnapPy census. The Alexander polynomial is ΔK(t)=1t+t4t5+t7t9+t10t13+t14\Delta_{K}(t)=1-t+t^{4}-t^{5}+t^{7}-t^{9}+t^{10}-t^{13}+t^{14}, so the formal semigroup is 𝒮K={0,4,7,8,10,11,12}14\mathcal{S}_{K}=\{0,4,7,8,10,11,12\}\cup\mathbb{Z}_{\geq 14}.

Figure 16 shows the graph of the gap function and its convex hull (we omit the details). It consists of branches of types (a), (b), (c), (f), (g) and (h) of Figure 4 from the left. Then there is no other gap function with the same convex hull.

Refer to caption
Figure 16. The graph of the gap function and its convex hull (broken line) of t09847.

Next, let KK be the hyperbolic knot v2871. The Alexander polynomial is 1t+t4t5+t7t8+t9t11+t12t15+t161-t+t^{4}-t^{5}+t^{7}-t^{8}+t^{9}-t^{11}+t^{12}-t^{15}+t^{16}, so the formal semigroup is {0,4,7,9,10,12,13,14}16\{0,4,7,9,10,12,13,14\}\cup\mathbb{Z}_{\geq 16} and the gap set is <0{1,2,3,5,6,8,11,15}\mathbb{Z}_{<0}\cup\{1,2,3,5,6,8,11,15\}. Figure 17 shows the graph of the gap function and its convex hull. In this case, the graph consists of branches of types (a), (b), (c), (e), (f), (g) and (h) of Figure 4 from the left. Again, there is no other gap function with the same convex hull. ∎

Refer to caption
Figure 17. The graph of the gap function and its convex hull (broken line) of v2871.

Acknowledgement

The author would like to thank Kouki Sato for valuable communication.

References

  • [1] S. Allen, Using secondary upsilon invariants to rule out stable equivalence of knot complexes, Algebr. Geom. Topol. 20 (2020), no. 1, 29–48.
  • [2] C. Anderson, K. Baker, X. Gao, M. Kegel, K. Le, K. Miller, S. Onaran, G. Sangston, S. Tripp, A. Wood and A. Wright, L–space knots with tunnel number >1>1 by experiment, preprint. arXiv:1909.00790
  • [3] K. Baker and M. Kegel, Census L–space knots are braid positive, except for one that is not, preprint. arXiv:2203.12013
  • [4] K. Baker, M. Kegel and D. McCoy, The search for alternating and quasi-alternating surgeries on asymmetric knots, in progress.
  • [5] K. Baker and K. Motegi, Seifert vs. slice genera of knots in twist families and a characterization of braid axes, Proc. Lond. Math. Soc. (3) 119 (2019), no. 6, 1493–1530.
  • [6] M. Borodzik and M. Hedden, The Υ\Upsilon function of L–space knots is a Legendre transform, Math. Proc. Cambridge Philos. Soc. 164 (2018), no. 3, 401–411.
  • [7] M. Borodzik and C. Livingston, Heegaard Floer homology and rational cuspidal curves, Forum Math. Sigma 2 (2014), Paper No. e28, 23 pp.
  • [8] S. Boyer, C. McA. Gordon and L. Watson, On L–spaces and left-orderable fundamental groups, Math. Ann. 356 (2013), no. 4, 1213–1245.
  • [9] M. Culler, N. Dunfield, M. Goemer and J. Weeks, SnapPy, a computer program for studying the topology and geometry of 33–manifolds, available at https://snappy.math.uic.edu
  • [10] N. Dunfield and J. Rasmussen, A unified Casson–Lin invariant for the real forms of SL(2), preprint. arXiv:2209.03382
  • [11] P. Feller, J. Park and A. Ray, On the Upsilon invariant and satellite knots, Math. Z. 292 (2019), no. 3-4, 1431–1452.
  • [12] R. H. Fox, Free differential calculus II, Ann. of Math. (2) 59 (1954) 196–210.
  • [13] H. Fujii, Geometric indices and the Alexander polynomial of a knot, Proc. Amer. Math. Soc. 124 (1996), no. 9, 2923–2933.
  • [14] M. Hedden, On knot Floer homology and cabling. II, Int. Math. Res. Not. IMRN 2009, no. 12, 2248–2274.
  • [15] J. Hom, A note on cabling and L–space surgeries, Algebr. Geom. Topol. 11 (2011), no. 1, 219–223.
  • [16] J. Hom, Satellite knots and L–space surgeries, Bull. Lond. Math. Soc. 48 (2016), no. 5, 771–778.
  • [17] J. Hom, A note on the concordance invariants epsilon and upsilon, Proc. Amer. Math. Soc. 144 (2016), no. 2, 897–902.
  • [18] K. Kodama, The software “KNOT”, a tool for knot theory, available at http://www.math.kobe-u.ac.jp/HOME/kodama/knot.html
  • [19] D. Krcatovich, The reduced knot Floer complex, Topology Appl. 194 (2015), 171–201.
  • [20] D. Krcatovich, A restriction on the Alexander polynomials of L–space knots, Pacific J. Math. 297 (2018), no. 1, 117–129.
  • [21] P. Lisca and G. Matić, Transverse contact structures on Seifert 33–manifolds, Algebr. Geom. Topol. 4 (2004), 1125–1144.
  • [22] P. Lisca and A. Stipsicz, Ozsváth-Szabó invariants and tight contact 33–manifolds. III, J. Symplectic Geom. 5 (2007), no. 4, 357–384.
  • [23] C. Livingston, Notes on the knot concordance invariant upsilon, Algebr. Geom. Topol. 17 (2017), no. 1, 111–130.
  • [24] J. M. Montesinos, Surgery on links and double branched covers of S3S^{3}, Knots, groups, and 33–manifolds (Papers dedicated to the memory of R. H. Fox), pp. 227–259. Ann. of Math. Studies, No. 84, Princeton Univ. Press, Princeton, N.J., 1975.
  • [25] H. Morton, The Alexander polynomial of a torus knot with twists, J. Knot Theory Ramifications 15 (2006), no. 8, 1037–1047.
  • [26] K. Motegi and K. Tohki, Hyperbolic L–space knots and exceptional Dehn surgeries, J. Knot Theory Ramifications 23 (2014), no. 14, 1450079, 13 pp.
  • [27] P. Ozsváth and Z. Szabó, On knot Floer homology and lens space surgeries, Topology 44 (2005), no. 6, 1281–1300.
  • [28] P. Ozsváth and Z. Szabó, On the Heegaard Floer homology of branched double-covers, Adv. Math. 194 (2005), no. 1, 1–33.
  • [29] P. Ozsváth and Z. Szabó, Knot Floer homology and rational surgeries, Algebr. Geom. Topol. 11 (2011), no. 1, 1–68.
  • [30] P. Ozsváth, A. Stipsicz and Z. Szabó, Concordance homomorphisms from knot Floer homology, Adv. Math. 315 (2017), 366–426.
  • [31] J. Rasmussen and S. Rasmussen, Floer simple manifolds and L–space intervals, Adv. Math. 322 (2017), 738–805.
  • [32] H. Schubert, Über eine numerische Knoteninvariante, Math. Z. 61 (1954), 245–288.
  • [33] J. Stallings, Constructions of fibred knots and links, Algebraic and geometric topology (Proc. Sympos. Pure Math., Stanford Univ., Stanford, Calif., 1976), Part 2, pp. 55–60, Proc. Sympos. Pure Math., XXXII, Amer. Math. Soc., Providence, R.I., 1978.
  • [34] A. Stoimenow, Realizing Alexander polynomials by hyperbolic links, Expo. Math. 28 (2010), no. 2, 133–178.
  • [35] M. Teragaito, Hyperbolic L–space knots and their formal semigroups, to appear in Internat. J. Math.
  • [36] G. Torres, On the Alexander polynomial, Ann. of Math. (2) 57 (1953) 57–89.
  • [37] S. Wang, Semigroups of L–space knots and nonalgebraic iterated torus knots, Math. Res. Lett. 25 (2018), no. 1, 335–346.
  • [38] S. Wang, A note on the concordance invariants Upsilon and phi, preprint. arXiv:2007.11511
  • [39] S. Wang, A further note on the concordance invariants epsilon and upsilon, Proc. Amer. Math. Soc. 148 (2020), no. 2, 893–899.
  • [40] S. Wang, A note on the concordance invariant epsilon, Q. J. Math. 73 (2022), no. 1, 333–347.
  • [41] L. Watson, A surgical perspective on quasi-alternating links, Low-dimensional and symplectic topology, 39–51, Proc. Sympos. Pure Math., 82, Amer. Math. Soc., Providence, RI, 2011.
  • [42] X. Xu, On the secondary Upsilon invariant, Algebr. Geom. Topol. 21 (2021), no. 4, 1661–1676.